• No results found

Bayesian inference and model choice for Holling’s disc equation: a case study on an insect predator prey system

N/A
N/A
Protected

Academic year: 2020

Share "Bayesian inference and model choice for Holling’s disc equation: a case study on an insect predator prey system"

Copied!
24
0
0

Loading.... (view fulltext now)

Full text

(1)

1

Bayesian Inference and Model Choice for Holling's Disc Equation: A Case Study

1

on an Insect Predator-Prey System

2

3

Short running title: Estimating functional response parameters 4

5

N. E. Papanikolaou1, 4, H. Williams2, N. Demiris3, S. P. Preston2, P. G. Milonas1 and 6

T. Kypraios2 7

8

1

Department of Entomology and Agricultural Zoology, Benaki Phytopathological

9

Institute, St. Delta 8, Kifissia, 14561, Greece

10

2

School of Mathematical Sciences, University of Nottingham, University Park, NG7

11

2RD, UK

12

3

Department of Statistics, Athens University of Economics and Business, Patission

13

76, Athens, 10434, Greece

14

4

Corresponding author: e-mail: nepapanikolaou@yahoo.gr

15

16

Keywords: Functional response, Markov Chain Monte Carlo, Bayes Factors, 17

Ordinary differential equation models, Coccinellidae.

(2)

2

Abstract: The dynamics of predator-prey systems relate strongly to the density 19

(in)dependent attributes of the predator's feeding rate, i.e. its functional response. The 20

outcome of functional response models is often used in theoretical or applied ecology 21

in order to extract information about the mechanisms associated with the feeding 22

behavior of predators. The focus of this study centres upon Holling's type II functional 23

response model, commonly known as the disc equation, which describes an inverse-24

density dependent mortality caused by a single predator to its prey. A common 25

method to provide inference on functional response data involves nonlinear least 26

squares optimization, assuming independent Gaussian errors, an assumption often 27

violated in practice due to the heteroscedasticity which is typically present in the data. 28

Moreover, as prey depletion is common in functional response experiments, the 29

differential form of disc equation ought to be used in principle. We introduce a related 30

statistical model and adopt a Bayesian approach for estimating parameters in ordinary 31

differential equation models. In addition, we explore model uncertainty via Bayes 32

factors. Our approach is illustrated via the analysis of several data sets concerning the 33

functional response of a widespread ladybird beetle (Propyleaquatuordecimpunctata

34

L.) to its prey (Aphis fabae Scopoli), predicting the efficiency of this predator on a 35

common and important aphid species. The results showed that the approach 36

developed in this study is towards a direction for accurate estimation of the 37

parameters that determine the shape of the functional response of a predator without 38

having to make unnecessary assumptions. The R (www.r-project.org) code for fitting 39

(3)

3

Introduction

41

The concept of functional response, a fundamental aspect of community 42

ecology, d\escribes the relationship between per capita predator consumption and prey 43

density (Solomon 1949). Holling (1959a) proposed various types of functional 44

response to provide a better understanding of the components of predator-prey 45

interactions; namely, a linear (type I), a decelerating (type II), and a sigmoid (type 46

III). In other words, the prey consumption is assumed to increase linearly with prey 47

density or increase asymptotically to a plateau under type I and type II respectively, 48

while in a type III functional response one assumes that the prey consumption is 49

supposed to be of a sigmoid form (S-shaped) as prey density increases. Although 50

more complex forms of the classical prey-dependent functional responses exist (see, 51

for example Jeschke et al. 2002), a significant amount of interest has been drawn to 52

Holling's type II and III functional responses because of their simplicity and 53

tractability, balancing between reality and feasibility (see, for example Englund et al. 54

2011). Holling's modelling approach for type II functional responses illustrates an 55

inverse-density-dependent prey mortality model which is common among invertebrate 56

predators (Hassell et al. 1977). Examining the workings of predator's individuals, 57

Holling (1959b) developed a mechanistic model to explain their feeding behaviour, 58

commonly known as the disc equation, which is an ordinary differential equation 59

(ODE) of the form: 60

 

N aT aN dt

t dN

h   

1 (1)

61

where N denotes the prey density, a the predator's attack rate, i.e. the per capita prey 62

mortality at low prey densities, and Th the handling time which reflects the time that a 63

predator spends on pursuing, subduing, eating and digesting its prey. Despite its 64

(4)

4

focus on the disc equation to describe predator's feeding behaviour, developing 66

several concepts in ecology theory or modelling predator-prey dynamics (see, for 67

example Beddington 1975, Englund et al. 2011, Jeschke et al. 2002, Okuyama 2012a). 68

Thus, it has become a baseline model in the sense of its determinant effect on much of 69

modern ecology theory (Englund et al. 2011). 70

Given the importance of the disc equation on natural ecosystems, a number of 71

early published papers investigated various statistical methods to infer the attack rate 72

(a) and the handling time (Th) from experimental data (see for example Fan and Petitt 73

1994, Livdahl 1979, Livdahl and Stiven 1983, Okuyama 2012b). One approach that 74

has been commonly used is to linearize the disc equation to enable estimation of a and 75

Th within the framework of linear regression models. Linearizing a non-linear model, 76

sometimes by making simplifying assumptions, is a method that has been attractive in 77

the literature due to its ease of implementation. In this particular case, this can be done 78

easily; setting N = N0 on the right hand side of (1), where N0 denotes the initial prey 79

density, an analytic expression of N(t) is available:

 

0 0 0

1 aT N t aN N

t N

h  

 . By

80

rearranging and taking the reciprocals, one can derive expressions of the least square 81

estimates of a and Th explicitly (i.e. without numerical optimization). Nevertheless, 82

such an approach relies on the assumption that the resource is not depleted during the 83

experimental progress. Whilst there are cases in which that assumption is not 84

unreasonable, such as in parasitoid-host systems, there are several cases where the 85

resources are depleted over time; for instance, in predator-prey systems. Therefore, in 86

such cases the differential form of the disc equation is to be preferred over its linear 87

approximation, or the random predator equation (Rogers 1972) which is the integrated 88

(5)

5

Another approach to estimate the parameters of the disc equation given 90

experimental data involves non-linear least squares optimization assuming identically 91

and independently distributed (additive) Gaussian errors. However, such an 92

assumption not only is likely to be violated by the heteroscedasticity which often 93

arises in functional response data (Trexler et al. 1988), but this particular error 94

distribution does not seem natural either, especially at early stages of the experiment 95

where the number of prey consumed is low. 96

An interesting approach to modelling predation in functional response was 97

developed by Fenlon and Faddy (2006) who studied two alternative model classes for 98

such systems, one using likelihood-based inference for a beta-binomial model 99

accounting for overdispersion and a counting-process-based framework. Although 100

there are similarities to our basic modelling framework there are also important 101

differences, namely we follow a distinct (Bayesian) approach to inference and model 102

selection and our computational framework does not resort to asymptotic normality. 103

In addition, our model differs in the way it accounts for density dependence. 104

The main aim of this paper is to introduce a hierarchical model which in 105

principle can incorporate any of Holling's various types of functional response and 106

accounts for heteroscedasticity. Also, in spite of numerical differential equation 107

solvers making it perfectly feasible to use richer ODE models, there is still a tendency 108

for researchers to use simpler models on grounds of convenience (e.g. the random 109

predator equation). Therefore, we aimed to show that it is perfectly feasible to work 110

with the richer models, providing clear statistical evidence of the benefit of doing so. 111

In addition, we illustrate how one can estimate the parameters of this model within a 112

Bayesian framework and select between competing models (hypotheses) given 113

(6)

6

analysis of eight data sets which involve the functional response of a predatory insect 115

to its prey. In particular, the ladybird beetle Propylea quatuordecimpunctata L. 116

(Coleoptera: Coccinellidae) and its essential prey Aphis fabae Scopoli (Hemiptera: 117

Aphididae) were used as case study organisms. Aphis fabae is well recognized as a 118

serious pest of cultivated plants worldwide (Blackman and Eastop 2000), where P.

119

quatuordecimpunctata is a widely distributed aphidophagous coccinellid (Hodek et al. 120

2012). As a thoroughly estimating of biological control agents' functional response is 121

of importance, with this application we provide a quantified analysis of the intake rate 122

of P. quatuordecimpunctata as a function of A. fabae density. 123

124

Materials and Methods

125

Data Collection and Experimental Conditions 126

An A. fabae colony originated from a stock colony at the Biological Control 127

Laboratory, Benaki Phytopathological Institute, reared on Vicia faba L. plants at 20 ± 128

1 °C, 65 ± 2% RH and a photoperiod of 16:8 L:D. Propylea quatuordecimpunctata

129

was collected from Zea mays L. plants infested with Rhopalosiphum maidis Fitch in 130

Arta County (Northwestern Greece). The coccinellid was reared in large cylindrical 131

Plexiglass cages (50 cm length 30 cm diameter) containing A. fabae prey on potted V.

132

faba plants at 25 ± 1 °C, 65 ± 2% RH and a photoperiod of 16:8 L:D. The 133

experiments were carried out at 20 ± 1 °C, 65 ± 2% RH and a photoperiod of 16:8 134

L:D. The experimental arena consisted of a plastic container (12cm height x 7cm 135

diameter) with a potted V. faba plant host (at 8-9cm height, top growth was cut) with 136

different A. fabae densities (3-3.5 days-old). An individual larva, female or male of P.

137

quatuordecimpunctata was placed into plastic containers, having starved for 12h. 138

(7)

7

16 and 32 aphids for 1st instar larvae, 2, 4, 8, 16, 32 and 64 aphids for 2nd instar larvae, 140

4, 8, 16, 32, 64 and 128 aphids for 3rd and 4th instar larvae as well as female and male 141

adults. We used 20-30 day old P. quatuordecimpunctata adults. Ten replicates of each 142

prey density were formed. Functional response experiments were also run at 25 ± 1 143

°C for female and male adults. The data sets concerning the functional response of 144

larvae were used in a previous study of Papanikolaou et al. (2011). 145

146

A Hierarchical Model 147

Denote by Ne

 

t the number of prey eaten by time t. Since a prey item is either 148

dead or alive by time t (which often denotes the end of the experiment), we assume 149

that Ne

 

t follows a Binomial distribution with parameters N0 and p

 

t , where N0 is

150

the initial prey population and p

 

t is the probability that a prey item has been eaten 151

by time t: 152

 

t

Ne Binom

N0,p

 

t

153

t a,T

N0 N

t a,T

/N0

p h   h (2)

154

where N

 

t is given by the solution of the ordinary differential equation (1) and 155

evaluated at time t. Notice that (1) cannot be solved analytically and hence the 156

solution has to be derived numerically. Furthermore, in principle any functional 157

response model for N

 

t can be used and not just the model as given in (1). 158

Bayesian Inference 159

Preliminaries 160

Traditionally, parameter estimation for models concerned with functional response 161

has been done by searching for the set of parameters (i.e. attack rate, handling time, 162

(8)

8

such as the sum of squared differences. Such an ordinary least squares (OLS) 164

approach provides an estimate for the parameter values that gives the “best fit” to the 165

experimental data, but it gives no information about uncertainty in the estimate; for 166

example, whether or not there are other plausible values of parameters that also give 167

equally good fits. Thus, being able to quantify the uncertainty surrounding the ability 168

of our point estimates to reflect the (unknown) truth is an equally important aspect in 169

parameter estimation. Typically, researchers resort to normality assumptions whence 170

OLS coincide with the maximum likelihood estimators (MLEs), leading to 171

quantification of the uncertainty around the MLEs. 172

In this paper we adopt the Bayesian paradigm which enables us to quantify the 173

uncertainty of our estimates in a coherent, probabilistic manner (e.g. Bolker 2008). 174

We utilise a Markov Chain Monte Carlo (MCMC) algorithm (see, for example, 175

Brooks et al. 2011) to sample from the posterior density of the parameters of interest 176

177

Likelihood, Prior and Posterior Distributions 178

Prior Distributions

179

We assume little prior knowledge of the attack rate (a) and handling time (Th) 180

when making inference for the parameters of our model. In particular, we assume that 181

both of them have independent slowly varying Exponential distributions: 182

a Exp

 

1 (3) 183

h

T Exp

 

2 (4) 184

and we typically set λ1 and λ2 to 10-6 in order to achieve large prior variance.

185

Assigning Exponential distribution with low rates is a typical choice when one is 186

interested in assuming a non-informative distribution about the parameters. In other 187

(9)

9

plausible positive real numbers (e.g. between 0 and 100), allowing for the data to 189

mostly inform the posterior density of a and Th. We have used non-informative priors 190

for the attack rate and the handling time. Although the maximum likelihood estimates 191

will coincide with the maximum a posteriori probability estimates in this case, we 192

advocate the use of a Bayesian approach since, in principle, one can assign 193

informative priors to either parameter (e.g. using information from past experiments) 194

and most importantly, offers a particularly natural way to select between candidate 195

models. 196

Likelihood

197

We now derive the likelihood of the observed data under the proposed 198

hierarchical model. Given that all the experiments lasted for 24 hours and for the ease 199

of exposition, we drop the dependence of t in the notation. Denote by X=

(

xi,ni

)

k

{

}

,

200

m

i1,...., and k 1,....,Kthe observed data of a functional response experiment; the 201

index i refers to the different initial prey densities that were used in the experiment 202

and k refers to each replication. Essentially, the observed data consist of pairs of 203

initial prey density and number of prey eaten after 24 hours. 204

An observed dataset is presented in Table 1 for illustration; the second column 205

nj

( )

consists of the initial prey densities for j1,....,6 and the rest,

(

x1,....,xj

)

refer 206

to the number of prey eaten by the predatorafter 24 hours. 207

The probability of observing xj prey items eaten out of njprey after 24 hours is 208

given by: 209

 

j x

n x

h j

j

j

p p x n T a n x x

P   

     

 0, , 1 (5)

(10)

10

where pP

t a,Th

is given in Equation 2 for t 24 and therefore is implicitly 211

dependent upon a and Th via (1). Assuming independence between the k replicates 212

in each experiment as well as between the different experiments, the likelihood of the 213

observed data X given the parameters  

a,Th

after T 24 hours is written as 214

follows: 215



 

j

h j j

h X a P x n a T

T a

L , , , , . (6)

216

Posterior Distribution

217

Equations 3, 4 and 6 give rise to the posterior distribution whose density is given as 218

follows: 219



 

k j

h h

j j

h X P x n a T a T

T

a, , , 12exp 12

 (7)

220

The posterior density of interest (Equation 7) is not of a closed form due to its 221

normalising constant not being available explicitly. Therefore, in this study we 222

employed to a random walk Metropolis algorithm (Gamerman and Lopes 2006) to 223

draw samples from 

a,Th X

.

224

Bayesian Model Choice

225

Statistical inference, in general, is not limited to parameter estimation. Another 226

common goal is hypothesis testing, in which we are interested in discriminating 227

models in order to gain a better understanding of the structure of the statistical 228

model(s) of interest and facilitate for model-robust decision making. Here we are 229

interested in observing the extent to which the observed data support the scientific 230

hypothesis that the differential form of the disc equation is to be used when prey is 231

depleted during the functional response experiments. 232

Bayes Factors

(11)

11

The Bayesian approach to model selection (or discrimination) is based upon an 234

extension to the posterior distribution to include not only uncertainty regarding the 235

model parameters but also for the model itself. Consider the following framework: 236

suppose we observe data X and have a series of plausible models indexed by 237

W

w1,...., . Denote by w the vector of parameters associated with model Mw and 238

by w

Xw

the likelihood of the observed data under model w. Then by specifying 239

a prior distribution pk

 

w for the model parameters under each model and a prior 240

probability for each model,p

 

Mw , we can derive the joint posterior distribution over 241

both the model and parameter spaces, given by 242

w Mw X

w

Xw

    

wwMw

 ,  (8)

243

Assuming prior independence between Mw and w, the joint posterior distribution 244

can then be written down (using Bayes Theorem) as product of two components: 245

w w X

w w X

 

w X

 ,  , (9)

246

where 

w,Mw X

is the posterior distribution of the parameters under model Mw

247

and 

Mw X

denotes what we refer to as the “posterior model probability” which

248

represents our beliefs, after observing data X , of what is the chance that model Mw

249

is the true model given that one of models 1,....,W is true. 250

Once these posterior model probabilities are obtained they can then be used to 251

discriminate between the competing models by computing the Bayes Factor which is 252

simply defined as the ratio of the posterior odds, i.e. the ratio of the posterior to the 253

prior model probability: 254

 

2

 

2

1 1

12

 

w w

w w

M M

X

M M

X BF

 

 

(10) 255

(12)

12

posterior odds = Bayes factor  prior odds. 257

The value of the Bayes factor represents the relative likelihood of M1 to M2 and is of 258

practical appeal because its value is independent of the choice of the prior model 259

probabilities (see Kass and Raftery 1995). It is easy to see that when the models are 260

equally probable a priori so that 

Mw 1



Mw 2

0.5the Bayes factor is 261

equal to the posterior odds in favour of M1. The quantity 

X Mw

for k 1,2 in 262

(10) is obtained by integrating over the parameter space, 263

  

w

w w w w w

w X M d

M X

   

 ,

264

where w is the parameter vector under model Mw and w

 

w is its prior density. 265

The term 

X Mw

is the marginal probability of the data and is often called the 266

marginal or integrated likelihood in the statistical literature while it is typically 267

referred to as the evidence in the physics and machine learning communities. The 268

Bayes factor is, therefore, a summary of the evidence provided by the data in favour 269

of one hypothesis represented by a statistical model as opposed to another. Note that 270

this formulation is completely general and does not require nested models, as is 271

typically the case with likelihood ratio tests. Additionally, no asymptotic justification 272

is required so that these results can be used for moderate sample sizes as well. 273

The marginal likelihoods are rarely available in analytic form. Therefore, in 274

practice if the number of parameters in each model is not very large (typically 2-5 275

parameters), then the marginal likelihoods and consequently the Bayes factors are 276

obtained via straightforward numerical integration. However, if the dimension of the 277

parameter vector θw is very large then computational tools such as trans dimensional

278

MCMC algorithms (Green 1995) can be used instead to explore the more complex 279

(13)

13 281

Results and Discussion

282

The functional response is a fundamental characteristic of predator-prey 283

systems. We have developed a hierarchical model which accounts for 284

heteroscedasticity and illustrated how to infer the parameters of interest (e.g. the 285

attack rate and the handling time) within a Bayesian framework using Markov Chain 286

Monte Carlo methods. In addition, we showed how one can assess competing 287

scientific hypotheses by investigating which model is mostly supported by the 288

experimental data. Generally, ODEs are frequently used in representing consumer-289

resource interactions and the outcome of such models is therefore of great interest to 290

researchers. Thus, we have made our computer code implementing the present 291

analysis in R (R Core team 2013) publicly available on http://www.maths.nott.ac.uk/ 292

~tk/files/functional_response/, to encourage and allow researchers to fit (and compare)

293

the proposed models to their datasets. 294

In practice, we often summarize the posterior distribution of the parameters by 295

calculating a variety of interpretable summary statistics such as posterior means, 296

medians and credible intervals. The posterior means of both parameters of the disc 297

equation obtained are presented in Table 2. By inspecting the 95% credible intervals 298

we observe that the estimated attack rates were similar for all four larval stages of the 299

predator, indicating that the larvae have similar abilities to respond to increasing prey 300

densities. On the other hand, handling times decreased for the older larvae. This 301

further indicates an increase in the upper level of the response, leading older larvae to 302

a higher consumption of prey. Being larger gives them an advantage in handling prey. 303

At 20 °C, the attack rate for females was higher than those for males. This means that 304

(14)

14

satiated) the females have the ability to consume more prey items than the males. 306

However, comparison of handling times yielded no differences, indicating that both 307

sexes have similar maximum predation ability. Overall, at 20 °C we expect that 308

females, males and fourth instar larvae of P. quatuordecimpunctata to display the 309

higher predation ability among predators stages. This could be of great interest for 310

biological control practitioners, since these stages are to be preferred in potential 311

release of this predator in agroecosystems, allowing an influential decrease of aphid 312

pests. 313

Our results also showed that at the temperature of 25 °C there was a notable 314

difference of estimated handling times between males and females. This further 315

indicates that females might prey and subdue prey more efficiently and faster than 316

males. Moreover, handling time increased considerably as temperature decreased 317

from 25 °C to 20 °C for females, but not for males. According to Papanikolaou et al. 318

(2013), the fecundity of P. quatuordecimpunctata females is higher at 25 °C than 20 319

°C, where females of roughly 20-30 day-old exhibit their maximum reproductive 320

potential at 25 °C. As a consequence, higher energy requirements for egg production 321

lead them to higher consumption of prey. Additionally, attack rate for males was 322

lower at 20 °C than 25 °C unlike females, as it was not different among these 323

temperatures. Attack rate might follow a hump-shaped relationship with temperature 324

as it happens for the ladybird Coleomegilla maculata lengi DeGeer (Sentis et al. 325

2012). The two temperatures examined here might have been at the plateau of the 326

hump-shaped relationship with temperature for females and therefore no differences 327

occurred, whereas, for males was still increasing with temperature. 328

Although investigating the Pearson’s correlation between the estimated 329

(15)

15

literature, it is important to do so since this may reveal potential parameter non-331

identifiability issues as well as biological insights. Table 2 reveals a moderate but 332

statistically significant positive correlation between the estimated handling times and 333

the estimated attack rates of the predator, based on 95% credible intervals. This is 334

biologically intuitive since coccinellids are being highly voracious, especially larvae 335

which consume more prey items than they need for their development (Hodek et al. 336

2012). This trend may lead to a gradual increase of the handling time, as the attack 337

rate increases. 338

In a previous study (Papanikolaou et. al. 2011) the authors fitted the non-differential 339

form of the disc equation using a non-linear least squares approach, in order to 340

provide inference for the functional response of P. quatuordecimpunctata larvae. The 341

values of attack rates are notably lower than those estimated in the present analysis, 342

indicating that linearisation may induce estimation bias. The attack rate coefficient 343

illustrates the per capita prey consumption at low prey densities, indicating the initial 344

slope of the functional response curve. A biased estimate of this parameter leads to 345

underestimation of prey consumption at the lower prey densities, in which the 346

handling time is not the limiting factor of the predation. In addition, a high value of 347

the attack rate coefficient shows that the predator may exhibit stronger density-348

dependent predation behavior. In contrary, the values of the larvae handling times are 349

close to those estimated in the present analysis. Handling time depicts a more 350

complex behavior which includes a number of distinguish predator activities, such as 351

pursuing, subduing, eating and digesting a prey item. 352

353

354

Model Selection

(16)

16

We applied the proposed method in two cases: 356

a) Our hypothesis is translated into two different models, describing type II functional 357

responses; in particular 358

M1:

 

0 0

1 aT N aN dt t dN h    359

M2:

 

N aT aN dt t dN h    1 360

Note that the model M1 uses the functional response used Papanikolaou et al. (2011) 361

while M2 uses the hierarchical model that is proposed in Material and Methods. 362

b) In this case, our aim was to distinguish between type II and type III functional 363

responses, which is of importance in functional response studies (Juliano 2001), i.e.: 364

M2:

 

N aT aN dt t dN h    1 365

M3:

 

2 2

1 aT N aN dt t dN h    , 366

where the model M3 describes type III functional responses. 367

In each cases, we assumed that both models are equally likely a priori and 368

consider Exponential prior distributions for both parameters, a Exp

 

 , Th Exp

 

 369

. It is well known that the Bayes factor can be sensitive to the choice of model 370

parameter's prior distributions. Therefore, we computed the Bayes factor for a range 371

of different values of , namely, 0.01, 0.1, 1, and 10. We first computed the log of 372

the marginal likelihoods for both models via numerical integration and then the Bayes 373

Factors of model M2 versus M1 in the first case and M2 versus M3 in the second case. 374

Table 3 and 4 shows the Bayes Factors of model M2 versus model M1 and M2versus 375

M3, respectively, for the different datasets and for different prior distributions. It is 376

(17)

17

M1 is to be preferred). Furthermore, the conclusions appear to be robust to the 378

different choice of . 379

Type II functional responses are frequent in nature, especially among 380

aphidophagous ladybirds (Hodek et al. 2012) and are typically descibed by Holling's 381

disc equation, one of the most commonly used models in ecology. Our study allowed 382

us to predict the efficiency of P. quatuordecimpunctata on a common and important 383

aphid species. Since biological control practitioners often rely on functional response 384

studies to design and use efficiently biocontrol agents, an accurate and non-biased 385

estimation of the functional response parameters is of crucial importance. The 386

approach developed here is towards that direction, for a more precise estimation of the 387

parameters that determine the shape of the functional response of a predator. Also, 388

functional response parameters of P. quatuordecimpunctata preying on A. fabae may 389

be incorporated in predator-prey models evaluating the population dynamics of the 390

study organisms. 391

From a statistical viewpoint routine Bayesian inference and model selection for 392

ODE-based models remains a challenge for a number of reasons which relate to the 393

need for solving the ODEs numerically. With respect to the former one may extend 394

our methods by utilising gradient-based information for the construction of efficient 395

MCMC proposals. The issue of model selection can be further explored by 396

methodology based upon thermodynamic integration (Friel and Pettitt 2008). Such an 397

approach is appealing in cases where numerical integration might be infeasible due to 398

the large number of parameters in the model, resulting in the evaluation of high-399

dimensional integrals. These are important directions for future research. 400

401

References

(18)

18

Beddington J. 1975. Mutual interference between parasites or predators and its effect 403

on searching efficiency. J. Anim. Ecol. 44(1):331-340. 404

Blackman, R.L. and V.F. Eastop. 2000. Aphids on the world's crops. An identification

405

and information guide. John Wiley & Sons. 406

Bolker, B. 2008. Ecological Models and Data in R,Princeton University Press. 407

Brooks, S., A. Gelman, G. Jones and X.L. Meng. 2011. Handbook of Markov Chain

408

Monte Carlo. Taylor & Francis US. 409

Englund, G., G. Ohlund, C.L. Hein and S. Diehl. 2011. Temperature dependence of 410

the functional response. Ecol. Lett. 14(9):914-921. 411

Fan, Y. and F.L. Petitt. 1994. Parameter estimation of the functional response. 412

Environ. Entomol. 23(4):785-794. 413

Fenlon & Faddy (2006) Modelling predation in functional response Ecol. Model. 198: 414

154-162. 415

Friel, N. and A.N. Pettitt. 2008. Marginal likelihood estimation via power posteriors. 416

J. R. Stat. Soc. B 70(3):589-607. 417

Gamerman, D. and H.F. Lopes. 2006. Markov Chain Monte Carlo: Stochastic

418

Simulation for Bayesian Inference. Chapman and Hall/CRC. 419

Gelman, A., G. Roberts and W. Gilks. 1996. Efficient metropolis jumping rules. In 420

Bernado, J.M. et al. (eds), BayesianStatistics, volume 5, page 599. OUP. 421

Gelman, A., W.R. Gilks and G. Roberts. 1997. Weak convergence and optimal 422

scaling of random walk Metropolis algorithms. Ann. Appl. Probab. 7(1):110-120. 423

Green, P.J. 1995. Reversible jump Markov chain Monte Carlo computation and 424

Bayesian model determination. Biometrika 82(4):711-732. 425

Hassell, M., J. Lawton and J. Beddington. 1977. Sigmoid functional responses by 426

(19)

19

Hastings, W. 1970. Monte Carlo samping methods using Markov chains and their 428

applications. Biometrika 57: 97-109. 429

Hodek, I., H.F. van Emden and A. Honěk. 2012. Ecology and Behaviour of the

430

Ladybird Beetles (Coccinellidae). Wiley-Blackwell. 431

Holling, C.S. 1959a. The components of predation as revealed by a study of small-432

mammal predation of the European pine sawfly. Can. Entomol. 91:293-320. 433

Holling, C.S. 1959b. Some characteristics of simple types of predation and parasitism. 434

Can. Entomol. 91:385-398. 435

Jeschke, J.M., M. Kopp and R. Tollrian. 2002. Predator functional responses: 436

discriminating between handling and digesting prey. Ecol. Monogr. 72(1):95-437

112. 438

Juliano, S.A. 2001. Nonlinear curve fitting: predation and functional response curves. 439

In: S.M. Scheiner and J. Gurevitch (eds), Design and analysis of ecological

440

experiments. Oxford University Press, Oxford, UK, p.p. 178-196. 441

Kass, R.E. and A.E. Raftery. 1995. Bayes factors. J. Am. Stat. Ass. 90: 773-795. 442

Livdahl, T.P. 1979. Evolution of handling time: the functional response of a predator 443

to the 450 density of sympatric and allopatric strains of prey. Evolution

444

33(2):765-768. 445

Livdahl, T.P. and A.E. Stiven. 1983. Statistical difficulties in the analysis of predator 446

functional response data. Can. Entomol. 115:1365-1370. 447

Metropolis, N., A.W. Rosenbluth, M.N. Rosenbluth, A.H. Teller and E. Teller. 1953. 448

Equation of State Calculations by Fast Computing Machines. J. Chem. Phys.

449

21:1087-1092. 450

Okuyama, T. 2012a. Flexible components of functional responses. J. Anim. Ecol.

451

(20)

20

Okuyama, T. 2012b. A likelihood approach for functional response models. Biol.

453

Contr. 60(2):103-107. 454

Papanikolaou, N.E., P.G. Milonas, D.C. Kontodimas, N. Demiris and Y.G. Matsinos. 455

2013. Temperature-dependent development, survival, longevity and fecundity of 456

Propyleaquatuordecimpunctata (Coleoptera: Coccinellidae). Ann. Entomol. Soc.

457

Am. 106(2):228-234. 458

Papanikolaou, N.E., A.F. Martinou, D.C. Kontodimas, Y.G. Matsinos and P.G. 459

Milonas. 2011. Functional responses of immature stages of Propylea

460

quatuordecimpunctata (Coleoptera: Coccinellidae) to Aphis fabae (Hemiptera: 461

Aphididae). Eur. J. Entomol. 108(3):391-395. 462

R Core Team. 2013. R: A Language and Environment for Statistical Computing.

463

Ripley, B.D. 1987. Stochastic simulation. John Wiley & Sons, Inc., New York, NY, 464

USA. 465

Rogers, D.J. 1972. Random search and insect population models. J. Anim. Ecol. 41: 466

369-383. 467

Sentis, A., J.L. Hemptinne and J. Brodeur. 2012. Using functional response modeling 468

to investigate the effect of temperature on predator feeding rate and energetic 469

efficiency. Oecologia 169(4):1117-1125. 470

Solomon, M. 1949. The natural control of animal populations. J. Anim. Ecol. 18(1):1-471

35. 472

Trexler, J., C. McCulloch and J. Travis. 1988. How can the functional response best 473

(21)

21

Table 1. Number of prey items consumed by Propylea quatuordecimpunctata

male adults for each trial (i=1,…,10). The experiment was conducted at 20 °C

for six different Aphis fabae prey densities (nij, j = 1, …, 6). Therefore, xij

denotes the count of consumed prey at the j-th density at the i-th trial.

j nj x1j x2j x3j x4j x5j x6j x7j x8j x9j x10j

1 4 4 4 4 4 4 4 4 4 4 4

2 8 8 8 8 8 8 2 8 8 7 7

3 16 8 14 10 10 15 14 12 14 9 16

4 32 16 27 18 16 23 20 17 21 31 17

5 64 30 29 33 24 30 22 20 26 26 27

6 128 50 36 28 26 24 41 30 38 28 42

[image:21.595.93.499.181.380.2]
(22)

22

Table 2. Parameter values of Holling’s disc equation obtained as posterior

means (95% Credible Intervals), and the correlation of attack rate and handling

time (95% Credible Intervals).

attack rate handling time correlation

1st instar

0.1496

(0.0728-0.2578)

7.1195

(5.6887-8.7337)

0.2393

(0.2197-0.2587)

2nd instar

0.1324

(0.0976-0.1763)

2.6713

(2.3351-3.0357)

0.1035

(0.0830-0.1239)

3rd instar

0.1514

(0.1230-0.1864)

1.1567

(1.0596-1.2605)

0.0453

(0.0246-0.0659)

4th instar

0.2025

(0.1744-0.2373)

0.5215

(0.4865-0.5575)

0.0864

(0.0659-0.1069)

females (20 °C)

0.2278

(0.1898-0.2737)

0.5058

(0.4071-0.6273)

0.0728

(0.0523-0.0934)

males (20 °C)

0.1067

(0.0889-0.1265)

0.6507

(0.5143-0.7735)

0.1600

(0.1396-0.1798)

females (25 °C)

0.2193

(0.1910-0.2494)

0.2565

(0.2237-0.2881)

0.1538

(0.1335-0.1740)

males (25 °C)

0.1970

(0.1666-0.2321)

0.4805

(0.4104-0.5608)

0.1994

(0.1795-0.2192)

[image:22.595.113.508.155.724.2]
(23)

23

Table 3. The Bayes Factor of M2 versus M1 for different values of the prior's

hyperparameter λ.

λ=0.01 λ=0.1 λ=1 λ=10

1st instar 8.17 7.61 4.04 0.27

2nd instar 175.91 170.71 151.21 8.50

3rd instar 1.65107 1.62107 1.31107 1.64106

4th instar 1.091021 1.061021 8.501020 9.211019

females (20 °C) 1.321020 1.291020 9.981019 7.951018

males (20 °C) 0.12 0.12 0.11 0.04

females (25 °C) 7.561022 7.411022 6.001022 7.811021

males (25 °C) 5.601014 5.491014 4.501014 5.561013

477

[image:23.595.98.512.127.435.2]
(24)

24

Table 4. The Bayes Factor of M2 versus M3 for different values of the prior's

hyperparameter λ.

λ=0.01 λ=0.1 λ=1 λ=10

1st instar 1.47 1.53 2.08 1.55

2nd instar 29.83 29.93 30.67 20.91

3rd instar 56.41 55.56 47.78 10.70

4th instar 1.69 1.67 1.43 0.30

females (20 °C) 132614 130522 111268 22586

males (20 °C) 2.641014 2.611014 2.331014 7.381013

females (25 °C) 614829 604344 508693 91812

males (25 °C) 6309 6230 5497 1559

[image:24.595.97.512.127.434.2]

Figure

Table 1. Number of prey items consumed by Propylea quatuordecimpunctata
Table 2. Parameter values of Holling’s disc equation obtained as posterior
Table 3. The Bayes Factor of M2 versus M1 for different values of the prior's
Table 4. The Bayes Factor of M2 versus M3 for different values of the prior's

References

Related documents

International Journal of Industrial Organization, Journal of Economics and Management Strategy, Journal of the European Economic Association, Journal of Labor Economics, Journal

The transportation capacities at the Entry Points, interconnected with foreign pipelines or LNG terminals, are calculated by means of hydraulic simulations of the

In conclusion, carbon dioxide functions as an accurate analog of nitrous oxide for self- pressurizing propellant tank testing at low to medium mass flow rates, but once the flow rates

• Analog to this an adaption of „ EEG- Umlage“ is advisably Spot price Energy price consumption <1 ct/kWh 2.5 ct/kWh Extra charge Energy output 1 bis 8 ct/kWh +3.5 ct/kWh

Unit costs at 2013/2014 prices 10 – 12 were applied to the resource use in the model to estimate the total NHS cost of patient management from the time a patient entered the data

common causes of bilateral leg edema are idiopathic edema (in young women) and chronic venous insufficiency (in.

In specific, CENTROTEC was able to consolidate and in some areas strengthen its market position in its largest segment Gas Flue Systems, for which it posted revenue of EUR

3. In a new game, the position of the dealer button will be determined by having the dealer deal one up card to each player. The first player to act will be the player